首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 300 毫秒
1.
Abstract Three kinds of trisaccharides were prepared by digesting fucoidan from the brown alga Kjellmaniella crassifolia, with the extracellular enzymes of the marine bacterium Fucobacter marina. Their structures were determined as Δ4,5GlcpUA1-2(L-Fucp(3-O-sulfate)α1-3)D-Manp, Δ4,5GlcpUA1-2(L-Fucp(3-O-sulfate)α1-3)D-Manp(6-O-sulfate), and Δ4,5GlcpUA1-2(L-Fucp(2,4-O-disulfate)α1-3)D-Manp(6-O-sulfate), which indicated the existence of a novel polysaccharide in the fucoidan and a novel glycosidase in the extracellular enzymes. In order to determine the complete structure of the polysaccharide and the reaction mechanism of the glycosidase, the fucoidan was partially hydrolyzed to obtain glucuronomannan, which is the putative backbone of the polysaccharide, and its sugar sequence was determined as (-4-D-GlcpUAβ1-2D-Manpα1-)n, which disclosed that the main structure of the polysaccharide is (-4-D-GlcpUAβ1-2(L-Fucp(3-O-sulfate)α1-3)D-Manpα1-)n. Consequently, the glycosidase was deduced to be an endo-α-D-mannosidase that eliminatively cleaves the α-D-mannosyl linkage between D-Manp and D-GlcpUA residues in the polysaccharide and produces the above trisaccharides. The novel polysaccharide and glycosidase were tentatively named as sulfated fucoglucuronomannan (SFGM) and SFGM lyase, respectively.  相似文献   

2.
Cell aggregation in the marine sponge Microciona prolifera is mediated by a multimillion molecular-mass aggregation factor, termed MAF. Earlier investigations revealed that the cell aggregation activity of MAF depends on two functional domains: (i) a Ca2+-independent cell-binding domain and (ii) a Ca2+-dependent proteoglycan self-interaction domain. Structural analysis of involved carbohydrate fragments of the proteoglycan in the self-association established a sulfated disaccharide β-d-GlcpNAc3S-(1→3)-α-l-Fucp and a pyruvated trisaccharide β-d-Galp4,6(R)Pyr-(1→4)-β-d-GlcpNAc-(1→3)-α-l-Fucp. Recent UV, SPR, and TEM studies, using BSA conjugates and gold nanoparticles of the synthetic sulfated disaccharide, clearly demonstrated self-recognition on the disaccharide level in the presence of Ca2+-ions. To determine binding forces of the carbohydrate–carbohydrate interactions for both synthetic MAF oligosaccharides, atomic force microscopy (AFM) studies were carried out. It turned out that, in the presence of Ca2+-ions, the force required to separate the tip and sample coated with a self-assembling monolayer of thiol-spacer-containing β-d-GlcpNAc-(1→3)-α-l-Fucp-(1→O)(CH2)3S(CH2)6S- was found to be quantized in integer multiples of 30 ± 6 pN. No binding was observed between the two monolayers in the absence of Ca2+-ions. Cd2+-ions could partially induce the self-interaction. In contrast, similar AFM experiments with thiol-spacer-containing β-d-Galp4,6(R)Pyr-(1→4)-β-d-GlcpNAc-(1→3)-α-l-Fucp-(1→O)(CH2)3S(CH2)6S- did not show a binding in the presence of Ca2+-ions. Also TEM experiments of gold nanoparticles coated with the pyruvated trisaccharide could not make visible aggregation in the presence of Ca2+-ions. It is suggested that the self-interaction between the sulfated disaccharide fragments is stronger than that between the pyruvated trisaccharide.  相似文献   

3.
A novel heterodimeric β-galactosidase with a molecular mass of 105 kDa was purified from crude cell extracts of the soil isolate Lactobacillus pentosus KUB-ST10-1 using ammonium sulphate fractionation followed by hydrophobic interaction and affinity chromatography. The electrophoretically homogenous enzyme has a specific activity of 97 UoNPG/mg protein. The Km, kcat and kcat/Km values for lactose and o-nitrophenyl-β-D-galactopyranoside (oNPG) were 38 mM, 20 s-1, 530 M-1·s-1 and 1.67 mM, 540 s-1, 325 000 M-1·s-1, respectively. The temperature optimum of β-galactosidase activity was 60–65°C for a 10-min assay, which is considerably higher than the values reported for other lactobacillal β-galactosidases. Mg2+ ions enhanced both activity and stability significantly. L. pentosus β-galactosidase was used for the production of prebiotic galacto-oligosaccharides (GOS) from lactose. A maximum yield of 31% GOS of total sugars was obtained at 78% lactose conversion. The enzyme showed a strong preference for the formation of β-(1→3) and β-(1→6) linkages, and the main transgalactosylation products identified were the disaccharides β-D-Galp-(1→6)-D -Glc, β-D-Galp-(1→3)-D -Glc, β-D -Galp-(1→6)-D -Gal, β-D -Galp-(1→3)-D -Gal, and the trisaccharides β-D -Galp-(1→3)-D -Lac, β-D -Galp-(1→6)-D -Lac.  相似文献   

4.
A xyloglucan-specific endo-1,4β-glucanase (XcXGHA) from Xanthomonas citri pv. mangiferaeindicae has been cloned, expressed in Escherichia coli, purified and characterised. The XcXGHA enzyme belongs to CAZy family GH74 and has catalytic site residues conserved with other xyloglucanases in this family. At its optimal reaction conditions, pH 7.0 and 40 °C, the enzyme has a k cat/K M value of 2.2?×?107 min?1 M?1 on a tamarind seed xyloglucan substrate. XcXGHA is relatively stable within a broad pH range (pH 4–9) and up to 50 °C (t 1/2, 50 °C of 74 min). XcXGHA is proven to be xyloglucan-specific, and a glycan microarray study verifies that XcXGHA catalyses cleavage of xyloglucan extracted from both monocot and dicot plant species. The enzyme catalyses hydrolysis of tamarind xyloglucan in a unique way by cleaving XXXG into XX and XG (X is xylosyl-substituted glucose; G is unsubstituted glucose), is able to degrade more complex xyloglucans and notably is able to cleave near more substituted xyloglucan motifs such as L [i.e. α-l-Fucp-(1?→?2)-β-d-Galp-(1?→?2)-α-d-Xylp-(1?→?6)-β-d-Glcp]. LC-MS/MS analysis of product profiles of tamarind xyloglucan which had been catalytically degraded by XcXGHA revealed that XcXGHA has specificity for X in subsite ?1. The 3D model suggests that XcXGHA consists of two seven-bladed β-propeller domains with the catalytic center formed by the interface of these two domains, which is conserved in xyloglucanases in the GH74 family. However, the XcXGHA has two amino acids (D264 and R472) that differ from the conserved residues of other GH74 xyloglucanases. These two amino acids were predicted to be located on the opposite side of the active site pocket, facing each other and forming a closing surface above the active site pocket. These two amino acids may contribute to the unique substrate specificity of the XcXGHA enzyme.  相似文献   

5.
Energetics of the catalysis of Class II α-mannosidase (E.C.3.2.1.24) from Aspergillus fischeri was studied. The enzyme showed Kcat/Km for Man (α1-3) Man, Man (α1-2) Man and Man (α1-6) Man as 7488, 5376 and 3690 M?1 min?1, respectively. The activation energy, Ea was 15.14, 47.43 and 71.21 kJ/mol for α1-3, α1-2 and α1-6 linked mannobioses, respectively, reflecting the energy barrier in the hydrolysis of latter two substrates. The enzyme showed Kcat/Km as 3.56 × 105 and 4.61 × 105 M?1 min?1 and Ea as 38.7 and 8.92 kJ/mol, towards pNPαMan and 4-MeUmbαMan, respectively. Binding of Swainsonine to the enzyme is stronger than that of 1-deoxymannojirimycin.  相似文献   

6.
The kinetics of the binding of cyanide to ferric chloroperoxidase have been studied at 25°C and ionic strength 0.11 M using a stopped-flow apparatus. The dissociation constant (KCN) of the peroxidase-cyanide complex and both forward (k+) and reverse (k?) rate constants are independent of the H+ concentration over the pH range 2.7 to 7.1. The values obtained are kcn = (9.5 ± 1.0) × 10-5 M, k+. = (5.2 ± 0.5) × 104 M?1 sec?1 and k- = (5.0± 1.4) sec-1. In the presence of 0 06 M potassium nitrate the affinity of cyanide for chloroperoxidase decreases due to the inhibition of the forward reaction. The dissociation rate is not affected. The nitrate anion exerts its influence by binding to a protonated form of the enzyme, whereas the cyanide binds to the unprotonated form. Binding of nitrate results in an apparent shift towards higher pKa values of the ionization of a crucial heme-linked acid group. Hence the influence of this group can be detected in the accessible pH range. Extrapolation to zero nitrate concentration yields a value of 3.1±0.3 for the pKa of the heme-linked acid group.  相似文献   

7.
A fucoidan-utilizing marine bacterium, Fucophilus fucoidanolyticus, was cultivated in medium containing fucoidan from Cladosiphon okamuranus. The C. okamuranus fucoidan was digested into oligosaccharides with the intracellular enzymes of F. fucoidanolyticus, and their structures were determined by nuclear magnetic resonance analyses. Some of their structures are represented by one general structural formula, (-3L-Fucp1-3L-Fucp(4-O-sulfate)1-3L-Fucp(4-O-sulfate)1-3(D-GlcpUA1-2)L-Fucp1)m-3L-Fucp1-3L-Fucp(4-O-sulfate)1-3L-Fucp(4-O-sulfate) 1-3L-Fucp (m = 0, 1, 2, or 3). We concluded that all oligosaccharides obtained were derived from a sulfated-fucose-containing polysaccharide of C. okamuranus, which has a repeating unit of (-3L-Fucp1-3L-Fucp(4-O-sulfate)1-3L-Fucp(4-O-sulfate)1-3(D-GlcpUA1-2)L-Fucp1-).  相似文献   

8.
Two trypsin inhibitors, LA-1 and LA-2, have been isolated from ridged gourd (Luffa acutangula Linn.) seeds and purified to homogeneity by gel filtration followed by ion-exchange chromatography. The isoelectric point is atpH 4.55 for LA-1 and atpH 5.85 for LA-2. The Stokes radius of each inhibitor is 11.4 å. The fluorescence emission spectrum of each inhibitor is similar to that of the free tyrosine. The biomolecular rate constant of acrylamide quenching is 1.0×109 M?1 sec?1 for LA-1 and 0.8 × 109 M?1 sec?1 for LA-2 and that of K2HPO4 quenching is 1.6×1011 M?1 sec?1 for LA-1 and 1.2×1011M?1 sec?1 for LA-2. Analysis of the circular dichroic spectra yields 40%α-helix and 60%Β-turn for La-1 and 45%α-helix and 55%Β-turn for LA-2. Inhibitors LA-1 and LA-2 consist of 28 and 29 amino acid residues, respectively. They lack threonine, alanine, valine, and tryptophan. Both inhibitors strongly inhibit trypsin by forming enzymeinhibitor complexes at a molar ratio of unity. A chemical modification study suggests the involvement of arginine of LA-1 and lysine of LA-2 in their reactive sites. The inhibitors are very similar in their amino acid sequences, and show sequence homology with other squash family inhibitors.  相似文献   

9.
The glycoside composition and sequence of an extracellular polysaccharide flocculant of Klebsiella pneumoniae H12 was analyzed. GC and HPLC analysis of the acid-hydrolysate identified its constituent monosaccharides as D-Glc, D-Man, D-Gal, and D-GlcA in an approximate molar ratio of 3.9:1.0:2.3:3.6. To analyze the glycoside sequence, the polysaccharide was partially hydrolyzed by acid and enzyme treatment. GC, HPLC, TLC, MALDI-TOF/MS, and 1H- and 13C- NMR spectroscopy characterized the obtained oligosaccharides.

The results clarified the partial structure of H12 polysaccharide as a linear polymer of a unit of pentasaccharide with a side chain of one D-GlcA to D-Glc moiety (see below). Although the existence of other sequences or other constituent glycosides could not be fully excluded, H12 polysaccharide must be a novel types as such a complicated unit for a polymer has not so far been reported. The partial structure of a H12 polysaccharide flocculant is also discussed in this report.

→4)- α-D-Glcp-(1→2)-α-D-Manp-(1→3)-4,6-Pyr-β-D- 3 Galp-(1→4)-β-D-Galp-(1→ ↓

1 β-D-GlcpA  相似文献   

10.
Dissociation and alkali complex formation equilibria of nitrilotris(methylenephosphonic acid) (NTMP, H6L) have been studied by dilatometric, potentiometric and 31P NMR-controlled titrations. Dilatometry indicated the formation of alkali complexes ML (M=Li, Na, K, Rb, Cs) at high pH with a stability decreasing from Li to Cs. An efficient combination of potentiometric and NMR methods confirmed two types of alkali metal complexes MHL and ML. Stability constants for the equilibria following M+ + HL5− ? MHL4− and M+ + L6− ? ML5−, respectively, were determined: logKNaHL=1.08(0.07), logKKHL=0.86(0.08), logKNaL=2.24(0.03). Systematic errors are introduced by using alkali metal hydroxides as titrants for routine potentiometric determinations of dissociation constants pKa5app and pKa6app. Correction formulae were derived to convert actual dissociation constants pKa into apparent dissociation constants pKaapp (or vice versa). The actual dissociation constants were found: pKa5(H2L4− ? H+ + HL5−)=7.47(0.03) and pKa6(HL5− ? H+ + L6−)=14.1(0.1). The anisotropy of 31P chemical shifts of salts MnH6 − nL (M=Li, Na, n=0-5) is more sensitive towards titration (n) than isotropic solution state chemical shifts.  相似文献   

11.
Since vitamin E deficiency is associated with increased susceptibility of erythrocytes to hemolysis, we investigated the presence of tocopherol binding sites in human red blood cells. Erythrocytes were found to have specific binding sites for d-α-[3H]tocopherol with properties of receptors. Kinetic studies of binding demonstrated two binding sites: one with high affinity (equilibrium association constant Ka = 2.6·107 M?1), low capacity (7600 sites/cell) and the second with low affinity (Ka = 1.24·106 M?1), high capacity (150000 sites/cell). These sites are at least partly protein in nature.  相似文献   

12.
Genes of β-mannosidase 97 kDa, GH family 2 (bMann9), β-mannanase 48 kDa, GH family 5 (bMan2), and α-galactosidase 60 kDa, GH family 27 (aGal1) encoding galactomannan-degrading glycoside hydrolases of Myceliophthora thermophila C1 were successfully cloned, and the recombinant enzymes were purified to homogeneity and characterized. bMann9 displays only exo-mannosidase activity, the K m and k cat values are 0.4 mM and 15 sec?1 for p-nitrophenyl-β-D-mannopyranoside, and the optimal pH and temperature are 5.3 and 40°C, respectively. bMann2 is active towards galac-tomannans (GM) of various structures. The K m and k cat values are 1.3 mg/ml and 67 sec?1 for GM carob, and the optimal pH and temperature are 5.2 and 69°C, respectively. aGal1 is active towards p-nitrophenyl-α-D-galactopyranoside (PNPG) as well as GM of various structures. The K m and k cat values are 0.08 mM and 35 sec?1 for PNPG, and the optimal pH and temperature are 5.0 and 60°C, respectively.  相似文献   

13.
Two new asterosaponins, diplasteriosides A and B, bearing the same β-D-Fucp-(1→2)-β-D-Galp-(1→4)-[β-D-Quip-(1→2)]-β-D-Quip-(1→3)-β-D-Quip-(1→ oligosaccharide chains linked to the C6 atom of the known genins, 3-O-sulfates of thornasterols A and B, respectively, were isolated from the Antarctic Diplasterias brucei starfish along with the previously known asteriidoside A. The structures of the new compounds were elucidated by two-dimensional NMR spectroscopy and mass spectrometry. Cytotoxicities of the isolated asterosaponins against the human colon cancer HCT-116, human breast cancer T-47D cell line, and human melanoma cancer RPMI-7951 cell lines were studied.  相似文献   

14.
The goal of this work is an examination of capillary exchange models as mathematical operators. The concentration function relations for the Krogh cylinder of a single capillary, basic to many organ models, are studied via the theory of operators on the Lebesgue normed spacesL p[0,∞], (1<-p<-∞). A discussion is included of theL p -normsvis-à-vis the coefficient of variation currently used in finding capillary parameters and evaluating parameter searches. The capillary model determines two operators on the space of locally integrable functions: O K (relating extravascular concentration to intravascular) and K a, k (relating intravascular concentration to input), wherek is the ratio of permeabilitysurface area (PS) to extravascular volume, and α is the ratio of PS to flow. These operators are shown to induce contractive (‖O K p <-1, ‖K a, k p <-1), isotone, linear operators onL p . The uniform convergence relation $$K_{a,k} = \mathop {\lim _{(p)} }\limits_{N \to \infty } \left( {\sum\limits_{n = 0}^N {P_n (a)O_k^n } } \right)$$ (as operators onL p) is derived, whereP n (a) is the Poisson probabilitye ?a a n /n!. For the important special cases ofp=∞, 1, 2 the norms are found (‖Ok‖=‖Ka,kp=1). Consideration is also given to the norms and operators when the functions involved are limited to a finite interval of time.  相似文献   

15.
The rates of deuterium exchange reactions of malondialdehyde (MDA) and deuterated malondialdehyde (MDAd) have been studied as a function of acidity and the content of dimethyl sulfoxide (DMSO) in binary mixtures with D2O . MDA incorporates deuterium from D2O solutions in a first-order reaction with a rate constant (kobs) that depends on the acid concentration. From this dependence, a catalytic constant, kcat, can be derived (kcatMDA = 2.25 × 105M?s?1). Similar kinetic behavior was found for MDAd in H2O solutions, and in this case, kcatMDA = 1.56 × 105M?1s?1. Results from reactions of MDA and MDAd in identical H2OD2O mixtures show that primary and secondary isotope effects are small (kH/kD = 1.13) and that solvent isotope effects cause most of the differences found between reactions in D2O and H2O. Reactions in binary DMSOd6D2O mixtures show a six-fold rate increase as the proportion of DMSOd6 increases from 50% to 90%. These results also illustrate the relatively high reactivity of MDA at pH values well above its pKa and the importance of medium composition on its reaction rate.  相似文献   

16.
Pyranose 2-oxidase, a homotetrameric FAD-flavoprotein from the basidiomycete Trametes multicolor, catalyzes regioselectively the oxidation of the 1→6 disaccharides allolactose [β- -Galp-(1→6)- -Glc], gentiobiose [β- -Glcp-(1→6)- -Glc], melibiose [α- -Galp-(1→6)- -Glc], and isomaltose [α- -Glcp-(1→6)- -Glc] at position C-2 of their reducing moiety. The resulting glycosyl -arabino-hexos-2-uloses can be reduced specifically at C-1 by NAD(P)H-dependent aldose reductase from the yeast Candida tenuis. By this novel, two-step redox isomerization process the four disaccharide substrates could be converted to the corresponding keto-disaccharides allolactulose [β- -Galp-(1→6)- -Fru], gentiobiulose [β- -Glcp-(1→6)- -Fru], melibiulose [α- -Galp-(1→6)- -Fru], and isomaltulose (palatinose, [α- -Glcp-(1→6)- -Fru]) in high yields. These products could find application in food technology as alternative sweeteners.  相似文献   

17.
The three-dimensional model of the CtCBM35 (Cthe 2811), i.e. the family 35 carbohydrate binding module (CBM) from the Clostridium thermocellum family 26 glycoside hydrolase (GH) β-mannanase, generated by Modeller9v8 displayed predominance of β-sheets arranged as β-sandwich fold. Multiple sequence alignment of CtCBM35 with other CBM35s showed a conserved signature sequence motif Trp-Gly-Tyr, which is probably a specific determinant for mannan binding. Cloned CtCBM35 from Clostridium thermocellum ATCC 27405 was a homogenous, soluble 16 kDa protein. Ligand binding analysis of CtCBM35 by affinity electrophoresis displayed higher binding affinity against konjac glucomannan (K a = 2.5 × 105 M?1) than carob galactomannan (K a = 1.4 × 105 M?1). The presence of Ca2+ ions imparted slightly higher binding affinity of CtCBM35 against carob galactomannan and konjac glucomannan than without Ca2+ ion additive. However, CtCBM35 exhibited a low ligand-binding affinity K a = 2.5 × 10?5 M?1 with insoluble ivory nut mannan. Ligand binding study by fluorescence spectroscopy showed K a against konjac glucomannan and carob galactomannan, 2.4 × 105 M?1 and 1.44 × 105 M?1, and ΔG of binding ?27.0 and ?25.0 kJ/mol, respectively, substantiating the findings of affinity electrophoresis. Ca2+ ions escalated the thermostability of CtCBM35 and its melting temperature was shifted to 70°C from initial 55°C. Therefore thermostable CtCBM35 targets more β-(1,4)-manno-configured ligands from plant cell wall hemicellulosic reservoir. Thus a non-catalytic CtCBM35 of multienzyme cellulosomal enzymes may gain interest in the biofuel and food industry in the form of released sugars by targeting plant cell wall polysaccharides.  相似文献   

18.
The optimization of bioreactor operations towards swainsonine production was performed using an artificial neural network coupled evolutionary program (EP)-based optimization algorithm fitted with experimental one-factor-at-a-time (OFAT) results. The effects of varying agitation (300–500 rpm) and aeration (0.5–2.0 vvm) rates for different incubation hours (72–108 h) were evaluated in bench top bioreactor. Prominent scale-up parameters, gassed power per unit volume (P g/V L, W/m3) and volumetric oxygen mass transfer coefficient (K L a, s?1) were correlated with optimized conditions. A maximum of 6.59 ± 0.10 μg/mL of swainsonine production was observed at 400 rpm-1.5 vvm at 84 h in OFAT experiments with corresponding P g/VL and K L a values of 91.66 W/m3 and 341.48 × 10?4 s?1, respectively. The EP optimization algorithm predicted a maximum of 10.08 μg/mL of swainsonine at 325.47 rpm, 1.99 vvm and 80.75 h against the experimental production of 7.93 ± 0.52 μg/mL at constant K L a (349.25 × 10?4 s?1) and significantly reduced P g/V L (33.33 W/m3) drawn by the impellers.  相似文献   

19.
The stimulation by Mg2+, Na+, K+, NH4 +, and ATP of (Na+, K+)-ATPase activity in a gill microsomal fraction from the freshwater prawn Macrobrachium rosenbergii was examined. Immunofluorescence labeling revealed that the (Na+, K+)-ATPase α-subunit is distributed predominantly within the intralamellar septum, while Western blotting revealed a single α-subunit isoform of about 108 kDa M r. Under saturating Mg2+, Na+, and K+ concentrations, the enzyme hydrolyzed ATP, obeying cooperative kinetics with V M = 115.0 ± 2.3 U mg?1, K 0.5 = 0.10 ± 0.01 mmol L?1. Stimulation by Na+ (V M = 110.0 ± 3.3 U mg?1, K 0.5 = 1.30 ± 0.03 mmol L?1), Mg2+ (V M = 115.0 ± 4.6 U mg?1, K 0.5 = 0.96 ± 0.03 mmol L?1), NH4 + (V M = 141.0 ± 5.6 U mg?1, K 0.5 = 1.90 ± 0.04 mmol L?1), and K+ (V M = 120.0 ± 2.4 U mg?1, K M = 2.74 ± 0.08 mmol L?1) followed single saturation curves and, except for K+, exhibited site–site interaction kinetics. Ouabain inhibited ATPase activity by around 73 % with K I = 12.4 ± 1.3 mol L?1. Complementary inhibition studies suggest the presence of F0F1–, Na+-, or K+-ATPases, but not V(H+)- or Ca2+-ATPases, in the gill microsomal preparation. K+ and NH4 + synergistically stimulated enzyme activity (≈25 %), suggesting that these ions bind to different sites on the molecule. We propose a mechanism for the stimulation by both NH4 +, and K+ of the gill enzyme.  相似文献   

20.
Ralstonia solanacearum lectin (RSL), that might be involved in phytopathogenicity, has been defined as lFuc?Man specific. However, the effects of polyvalency of glycotopes and mammalian structural units on binding have not been established. In this study, recognition factors of RSL were comprehensively examined with natural multivalent glycotopes and monomeric ligands using enzyme linked lectin-sorbent and inhibition assays. Among the glycans tested, RSL reacted strongly with multivalent blood group Ah (GalNAcα1–3[Fucα1–2]Gal) and H (Fucα1–2Gal) active glycotopes, followed by Bh (Galα1–3[Fucα1–2]Gal), Lea (Galβ1–3[Fucα1–4]GlcNAc) and Leb (Fucα1–2Galβ1–3[Fucα1–4]GlcNAc) active glycotopes. But weak or negligible binding was observed for blood group precursors having Galβ1–3/4GlcNAcβ1- (Iβ/IIβ) residues or Galβ1–3GalNAcα1- (Tα), GalNAcα1-Ser/Thr (Tn) bearing glycoproteins. These results indicate that the density and degree of exposure of multivalent ligands of α1–2 linked lFuc to Gal at the non-reducing end is the most critical factor for binding. An inhibition study with monomeric ligands revealed that the combining site of RSL should be of a groove type to fit trisaccharide binding with highest complementarity to blood group H trisaccharide (HL; Fucα1–2Galβ1–4Glc). The outstandingly broad RSL saccharide-binding profile might be related to the unusually wide spectrum of plants that suffer from R. solanacearum pathogenicity and provide ideas for protective antiadhesion strategies.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号